• Aucun résultat trouvé

On the existence of L2-valued thermodynamic entropy solutions for a hyperbolic system with boundary conditions

N/A
N/A
Protected

Academic year: 2021

Partager "On the existence of L2-valued thermodynamic entropy solutions for a hyperbolic system with boundary conditions"

Copied!
14
0
0

Texte intégral

(1)

On the existence of L

2

-valued thermodynamic entropy solutions

for a hyperbolic system with boundary conditions

Stefano Marchesani

Stefano Olla

May 13, 2019

Abstract

We prove existence of L2-weak solutions of a quasilinear wave equation with boundary conditions.

This describes the isothermal evolution of a one dimensional non-linear elastic material, attached to a fixed point on one side and subject to a force (tension) applied to the other side. The L2-valued

solutions appear naturally when studying the hydrodynamic limit from a microscopic dynamics of a chain of anharmonic springs connected to a thermal bath. The proof of the existence is done using a vanishing viscosity approximation with extra Neumann boundary conditions added. In this setting we obtain a uniform a priori estimate in L2, allowing us to use L2 Young measures, together with the classical tools of compensated compactness. We then prove that the viscous solutions converge to weak solutions of the quasilinear wave equation strongly in Lp, for any p ∈ [1, 2), that satisfy, in a weak sense, the boundary conditions. Furthermore, these solutions satisfy the Clausius inequality: the change of the free energy is bounded by the work done by the boundary tension. In this sense they are the correct thermodynamic solutions, and we conjecture their uniqueness.

Keywords: hyperbolic conservation laws, quasi-linear wave equation, boundary conditions, weak solutions, vanishing viscosity, compensated compactness, entropy solutions, Clausius inequality. Mathematics Subject Classification numbers: 35L40, 35D40

1

Introduction

The problem of existence of weak solutions for hyperbolic systems of conservation law in a bounded domain has been studied for solutions that are of bounded variation or in L∞ [7]. In the scalar

case some works extend to L∞ solutions, obtained from viscous approximations [20]. But viscous

approximations require extra boundary conditions, that are usually taken of Dirichlet type. We present here an approach based on viscosity approximations, where the extra boundary conditions are of Neumann type, to reflect the conservative nature of the viscous approximation. We consider here the quasilinear wave equation

(

rt− px= 0

pt− τ (r)x= 0

, (t, x) ∈ R+× [0, 1] (1.1)

where τ (r) is a strictly increasing regular function of r such that 0 < c1 ≤ τ′(r) ≤ c2, for some

constant c1, c2. In section 2 we will require some more technical assumption for τ . We add to the

system the following boundary conditions:

p(t, 0) = 0, τ (r(t, 1)) = ¯τ (t) (1.2)

(2)

and initial data

r(0, x) = r0(x), p(0, x) = p0(x). (1.3)

The boundary tension ¯τ : R+→ R is smooth and bounded with bounded derivative.

The equations (1.1) describe the isothermal evolution of an elastic material in Lagrangian co-ordinates. The material point x ∈ [0, 1] has a volume strain r(t, x) at time t (that can also have negative values), and momentum (velocity) p(t, x). The Eulerian position of the material point x, with respect to the position of the particle 0, is given by q(t, x) = R0xr(t, y)dy, so that we can identify the position of the material point x = 1 as the total extension of the material:

L(t) = q(t, 1) = Z1

0

r(t, y)dy. (1.4)

Let T < ∞ be given and arbitrary, and define QT := [0, T ] × [0, 1]. We shall construct weak

solutions ¯u(t, y) = (¯r(t, y), ¯p(t, y)) , (t, y) ∈ QT, to the quasilinear wave equation such that ¯u(t, ·) ∈

L2(0, 1) for all t ≤ T and satisfy the initial and boundary conditions in the following weak sense:

Z 1 0 ϕ(t, x)¯r(t, x)dx − Z 1 0 ϕ(0, x)r0(x)dx = Z t 0 Z 1 0 (ϕsr − ϕ¯ xp) dxds¯ (1.5) Z 1 0 ψ(t, x)¯p(t, x)dx − Z 1 0 ψ(0, x)p0(x)dx = Z t 0 Z 1 0 (ψsp − ψ¯ xτ (¯r)) dxds + Zt 0 ψ(s, 1)¯τ (s)ds (1.6)

for all functions ϕ, ψ ∈ C1(Q

T) such that ϕ(t, 1) = ψ(t, 0) = 0 for all t ≥ 0.

Define the free energy of the system, associated to a profile u(x) = (r(x), p(x)) ∈ L2(0, 1), as

F(u) := Z 1 0  p2(x) 2 + F (r(x))  dx (1.7)

where F (r) is a primitive of τ (r) (F′(r) = τ (r)), such that c1

2r

2≤ F (r) ≤ c2

2r

2for any r ∈ R. This

is possible thanks to the bounds we required on τ′.

The solution ¯u of (1.6) that we obtain has the following properties: • ¯u ∈ L∞(0, T ; L2(0, 1))

• ¯u(0, x) = u0(x) for a.e. x;

• For any φ ∈ C1([0, 1]), the application t 7→

Z 1 0

φ(x)¯u(t, x)dx (1.8)

is Lipschitz continuous over [0, T ]; • ¯u satisfies Clausius inequality:

F(¯u(t)) − F(u0) ≤ W (t), ∀t ∈ [0, T ] (1.9) where u0= (r0, p0) and W (t) := − Z t 0 ¯ τ′(s) Z 1 0 ¯ r(s, x)dxds + ¯τ (t) Z 1 0 ¯ r(t, x)dx − ¯τ(0) Z1 0 r0(x)dx (1.10)

is the work done by the external tension up to time t. In this sense we call our solution a ther-modynamic entropy solution. For general discussion of the connection of such therther-modynamic solutions to the usual definition of entropic solutions, see [11] and [4].

Remark. If ¯r(t, x) is differentiable with respect to time, we may perform an integration by parts and obtain W (t) = Z t 0 ¯ τ (s)dL(s). (1.11)

(3)

The construction of the solution is obtained from the following viscosity approximation ( rtδ− pδx= δrδxx pδ t− τ (rδ)x= δpδxx , (t, x) ∈ R+× [0, 1] (1.12)

with boundary conditions

pδ(t, 0) = 0, τ (rδ(t, 1)) = ¯τ (t), pδx(t, 1) = 0, rxδ(t, 0) = 0 (1.13)

and initial data

rδ(0, x) = rδ0(x), pδ(0, x) = pδ0(x) (1.14)

such that rδ

0 and pδ0 are compatible with the boundary conditions, regular enough (see (3.4) and

(3.5)) and converge to r0 and p0, respectively, as δ → 0.

Note that in the viscous approximation we have added two Neumann boundary conditions, that reflect the conservative nature of the viscous perturbation. Under these conditions we have

Z 1 0 |uδ(t, x)|2dx + δ Z t 0 Z1 0 |uδx(s, x)|2dxds ≤ C, ∀t ≥ 0 (1.15)

where C is independent of t and δ. It is thus clear that {uδ}

δ>0 and {

√ δuδ

x}δ>0 are uniformly

bounded in L2(Q

T). Then we rely on the existence of a family of bounded Lax entropy-entropy

fluxes as in [23], [21] and [22], that allows us to apply the compensated compactness in the L2

version. The conditions assumed on τ (r) are in fact those required to apply [23] results. Under a slight different set of conditions, another Lp extension of the compensated compactness argument can be found in [13].

1.1

Physical motivations

The problem arises naturally considering hydrodynamic limit for a non-linear chain of anharmonic oscillators in contact with a heat bath at a given temperature [16, 15]. This microscopic dynamics models an isothermal transformation with two locally conserved quantities that evolve, on the macroscopic scale, following (1.1).

Consider N + 1 particles on the real line and, for i = 0, . . . N , call qiand pithe positions and the

momenta of the i-th particle, respectively. Particles i and i − 1 interact via a nonlinear potential V (qi− qi−1). Particle i = 0 is at position q0= 0 and does not move, i.e. p0(t) = 0. There is a time

dependent force (tension) ¯τ (t) acting on the last particle. Then, defining ri:= qi− qi−1 we have a

system with Hamiltonian

HN(t) = N X i=1  p2 i 2 + V (ri)  − ¯τ(t) N X i=1 ri. (1.16)

The interaction with a heat bath at temperature β−1 is modeled by a stochastic perturbation of the dynamics, that acts as a microscopic stochastic viscosity. Defining the discrete gradient and laplacian as

∇ai= ai+1− ai, ∆ai= ai+1+ ai−1− 2ai,

the evolution equations are then given by the following system of stochastic differential equations:                    dr1= p1dt + δ∇V′(r1)dt − p 2β−1δ d ew 1 dri= ∇pi−1dt + δ∆V′(ri)dt − p 2β−1δ ∇d ew i−1, 2 ≤ i ≤ N − 1 drN= ∇pN −1dt + δ (¯τ (t) + V′(rN −1) − 2V′(rN)) − p 2β−1δ ∇d ewN −1, dp1= ∇V′(r1)dt + δ (p2− 2p1) dt − p 2β−1δ ∇dw1, dpj= ∇V′(rj)dt + δ∆pjdt − p 2β−1δ ∇dw j−1, 2 ≤ j ≤ N − 1 dpN= (¯τ (t) − V′(rN))dt − δ∇pN −1dt + p 2β−1δ dwN −1 (1.17)

(4)

Here β−1> 0 is the temperature of the heat bath, and {wi}N −1i=1 , { ewi}N −1i=1 are families of independent

Brownian motions. The parameter δ is the intensity of the action of the heat bath, and is chosen depending on N such that δ ∼ o(N). When δ = 0, equations (1.17) are just the Newton deterministic equations for the Hamiltonian (1.16). Notice the correspondence of the boundary conditions in (1.17) with the one chosen in (1.12).

One of the effects of the action of the stochastic heat bath is to fix, in a large time scale, the variance of the velocities (i.e. the temperature) at β−1, and establish a local equilibrium, where

space-time averages of V′(r

i) around a macroscopic particle number [N x] at a macroscopic time

N t converges to the equilibrium tension τ (r(t, x), β) at temperature β−1and volume stretch r(t, x).

Since β is fixed by the heat bath and do not evolve in time, we drop it from the notation in the sequel.

The hydrodynamic limit consists in proving that, for any continuous function G(x) on [0, 1],

1 N N X i=1 G  i N   ri(N t) pi(N t)  −→ N →∞ Z 1 0 G(x)  r(t, x) p(t, x)  dx, (1.18)

in probability, with (r(t, x), p(t, x)) satisfying (1.5), (1.6). Of course a complete proof would require the uniqueness of such L2 valued solutions that satisfy (1.9): this remains an open problem. The

results contained in [15] states that the limit distribution of the empirical distribution defined on the RHS of (1.18), concentrates on the possible solutions of (1.5) and (1.6) that satisfy (1.9). Since we have no uniqueness result, we cannot assure that the solutions constructed in the present paper coincide with those obtained with the hydrodynamic limit from (1.17). One can however conjecture that this is the case.

This stochastic model was already considered by Fritz [12] in the infinite volume without bound-ary conditions, and in [16], but without the characterisation of the boundbound-ary conditions.

In the hydrodynamic limit only L2 bounds are available and we are constrained to consider L2 valued solutions. Since these solutions do not have definite values on the boundary, boundary

conditions have only a dynamical meaning in the sense of an evolution in L2given by (1.5), (1.6).

2

Hyperbolic system and the existence of weak solutions

For r, p : R+× [0, 1] → R, consider the hyperbolic system

( rt− px= 0 pt− τ (r)x= 0 , p(t, 0) = 0 r(t, 1) = τ −1τ (t)) p(0, x) = p0(x) r(0, x) = r0(x) (2.1)

The nonlinearity τ ∈ C3(R) is chosen to have the following properties.

(τ -i) c1≤ τ′(r) ≤ c2 for some c1, c2> 0 and all r ∈ R;

(τ -ii) τ′′

(r) 6= 0 for all r ∈ R; (τ -iii) τ′′(r), τ′′′

(r) ∈ L2(R) ∩ L

(R).

We also assume that ¯τ : R+→ R is smooth. Moreover, there is a time T⋆such that ¯τ′(t) = 0 for all

t ≥ T⋆. The initial data r0, p0∈ L2(0, 1) are compatible with the boundary conditions.

Remark. Conditions (τ -i) and (τ -ii) ensure that the system is strictly hyperbolic and genuinely nonlinear, respectively. Condition (τ -iii) is used later on to ensure some boundedness properties of the Lax entropies.

Theorem 2.1. System (2.1) admits a weak solution ¯u = (¯r, ¯p) in the sense of (1.5) and (1.6), such that ¯u ∈ L∞(0, T ; L2(0, 1)), ¯u(0, x) = u

0(x) for a.e. x; and it satisfies the Clausius inequality:

F(¯u(t)) − F(u0) ≤ W (t), ∀t ∈ [0, T ] (2.2)

(5)

3

Viscous approximation and energy estimates

We consider the following parabolic approximation of the hyperbolic system (2.1) ( rδ t − pδx= δrδxx pδ t− τ (rδ)x= δpδxx , (t, x) ∈ R+× [0, 1] (3.1)

for δ > 0, with the boundary conditions:

pδ(t, 0) = 0, rδ(t, 1) = τ−1(¯τ (t)), pδx(t, 1) = 0, rδx(t, 0) = 0, (3.2)

and initial data:

pδ(0, x) = pδ0(x), rδ(0, x) = rδ0(x). (3.3)

The initial data rδ

0, pδ0 ∈ C∞([0, 1]) are mollifications of r0 and p0 compatible with the boundary

conditions:

pδ0(0) = 0 rδ0(1) = τ−1(¯τ (0)), ∂xpδ0(1) = 0 ∂xrδ0(0) = 0. (3.4)

Moreover, there is C independent of δ such that

krδ0kL2+ kpδ0kL2+ k √ δ∂xr0δkL2. + k √ δ∂xpδ0kL2≤ C (3.5) and (rδ 0, pδ0) → (r0, p0) strongly in L2(0, 1).

As shown in [2] in a more general setting, this system admits a global classical solution (rδ, pδ),

with

rδ, pδ∈ C1(R+; C0([0, 1])) ∩ C0(R+; C2([0, 1])).

Remark. (i) We added two extra Neumann conditions, namely pδ

x(t, 1) = rδx(t, 0) = 0. These

conditions reflect the conservative nature of the viscous perturbation, and are required in order to obtain the correct production of free energy.

(ii) One could introduce a nonlinear viscosity term: δτ (rδ)

xx in place of δrxxδ . This is a term

which comes naturally from a microscopic derivation of system (3.1), as described in the introduction (see also [14]). Nevertheless, this does not drastically change the problem, thus we shall consider only the linear viscosity δrδxx.

Theorem 3.1(Energy estimate). There there is a constant C > 0 independent of t and δ such that Z 1 0 |uδ(t, x)|2dx + δ Z t 0 Z 1 0 |uδx(s, x)|2dxds ≤ C (3.6)

for all t ≥ 0 and δ > 0.

Proof. Let F be a primitive of τ such that c1 2r

2

≤ F (r) ≤ c22r2. By a direct calculation we have Z 1 0  (pδ)2 2 + F (r δ)  dx t=T t=0 + Z T 0 Z 1 0  δ(rδx)2+ δ(pδx)2  dxdt = Z T 0 ¯ τ (t) Z 1 0 rδtdxdt (3.7) =  ¯ τ (t) Z 1 0 rδdx t=T t=0 − Z T 0 ¯ τ′(t) Z1 0 rδdxdt. (3.8)

Write, for some ε > 0 to be chosen later,

¯ τ (T ) Z 1 0 rδ(T, x)dx ≤ |¯τ(T )|  1 2ε+ ε 2 Z 1 0 (rδ)2(T, x)dx  ≤C¯τ +Cτ¯ε 2 Z1 0 (rδ)2(T, x)dx (3.9) where Cτ¯= supt≥0(|¯τ(t)| + |¯τ ′ (t)|) depends on ¯τ only.

(6)

Using F (r) ≥c21r2 we obtain  c1 2 − Cτ¯ε 2  Z 1 0 (rδ)2(T, x)dx +1 2 Z 1 0 (pδ)2(T, x)dx + Z T 0 Z 1 0  δ(rδx)2+ δ(pδx)2  dxdt ≤ Cτ¯ 2ε + C0− Z T 0 ¯ τ′(t) Z 1 0 rδ(t, x)dxdt. (3.10)

Recall that there is T⋆> 0 such that ¯τ′(t) = 0 for t ≥ T⋆. Then, for T < T⋆, we write

 c1 2 − Cτ¯ε 2  Z 1 0 (rδ)2(T, x)dx +1 2 Z 1 0 (pδ)2(T, x)dx + Z T 0 Z 1 0  δ(rδx)2+ δ(pδx)2  dxdt ≤ C¯τ + C0+ C2 ¯ τ 2 T + 1 2 ZT 0 Z1 0 (rδ)2(t, x)dx (3.11) ≤Cτ¯ + C0+C 2 ¯ τ 2 T + 1 2 Z T 0 Z 1 0  (rδ)2(t, x) + (pδ)2(t, x)dxdt (3.12) +1 2 Z T 0 Z t 0 Z 1 0  δ(rδx)2+ δ(pδx)2  dxdsdt

where C0 depends on the initial data only. Choosing ε = c1/(2Cτ¯) gives

c1 4J(T ) ≤ C0+ C2¯τ c1 +C 2 ¯ τ 2 T + 1 2 Z T 0 J(t)dt, (3.13) where J(t) = Z 1 0  (rδ)2(t, x) + (pδ)2(t, x)dx + Z t 0 Z 1 0  δ(rxδ)2+ δ(pδx)2  dxds. (3.14)

We apply Gronwall’s inequality. This, together with T < T⋆, gives

J(T ) ≤4c1C0+ 2C 2 ¯ τ(2 + c1T ) c2 1 exp  2T c1  ≤4c1C0+ 2C 2 ¯ τ(2 + c1T⋆) c2 1 exp  2T⋆ c1  := C0(c1, ¯τ ), (3.15)

for all T ∈ [0, T⋆), where C0(c1, ¯τ ) is independent of T and δ

On the other hand, if T ≥ T⋆, we have

 c1 2 − Cτ¯ε 2  Z 1 0 (rδ)2(T, x)dx +1 2 Z 1 0 (pδ)2(T, x)dx + Z T 0 Z 1 0  δ(rδx)2+ δ(pδx)2  dxdt ≤ C¯τ + C0− Z T⋆ 0 ¯ τ′(t) Z 1 0 rδ(t, x)dxdt (3.16)

and the integral at the right-hand side is uniformly bounded in T , δ and δ, since

− ZT⋆ 0 ¯ τ′(t) Z 1 0 rδ(t, x)dxdt ≤ Cτ¯ ZT⋆ 0 Z 1 0 |rδ(t, x)|dxdt ≤ Cτ¯T⋆  1 T⋆ Z T⋆ 0 Z 1 0 (rδ)2(t, x)dxdt 1/2 ≤ Cτ¯ √ T⋆ Z T⋆ 0 J(t)dt 1/2 ≤ Cτ¯T⋆ p C0(c1, ¯τ ) (3.17)

(7)

From (3.7) we also immediately obtain the following Corollary 3.2 (Viscous Clausius inequality).

F(uδ(t)) − F(uδ0) ≤ − Z t 0 ¯ τ′(s) Z 1 0 rδ(s, x)dx + ¯τ (t) Z 1 0 rδ(t, x)dx − ¯τ(0) Z1 0 r0δ(x)dx. (3.18)

4

L

2

Young measures and compensated compactness

Throughout this section, for any fixed T > 0 let uδ(t, x) := (rδ(t, x), pδ(t, x)) be a strong solution

of (3.1) on QT. By Theorem 3.1 and after a time integration over [0, T ] we obtain

kuδkL2(QT)≤ C (4.1)

for some C independent of δ. Thus we can extract from {uδ}

δ>0 a subsequence that is weakly

convergent in L2(Q

T). Namely, up to a subsequence, there exists ¯u = (¯r, ¯p) ∈ L2(QT) such that

lim δ→0 Z QT uδϕ = Z QT ¯ uϕ, ∀ϕ ∈ L2(QT). (4.2)

All the limits δ → 0 taken below are intended along a chosen subsequence. In this section we want to show that for any φ ∈ L2(QT) we have

lim δ→0 Z QT φ(t, x)τ (rδ(t, x)) dx dt = Z QT φ(t, x)τ (¯r(t, x)) dx dt. (4.3)

This is done using a L2version of the compensated compactness, which is usually performed in L∞.

From the solution uδ(t, x), we define the following Young measure on Q T× R2:

νt,xδ := δuδ(t,x), (4.4)

which is a Dirac mass centred at uδ, i.e.

Z QT J(t, x)f (uδ(t, x)) dx dt = Z QT Z R2 J(t, x)f (ξ)dνt,xδ (ξ)dxdt

for all mesurable J : QT → R and f : R2→ R.

Since we have L2 bounds on uδ, we refer at νδ

t,x as a L2-Young measure [3]. In particular we

have, from (4.1) Z QT Z R2|ξ| 2 dνt,xδ (ξ) dx dt ≤ C. (4.5)

We call Y the set of Young measures on QT× R2 and we make it a metric space by endowing it

with the Prohorov’s metric. By proposition 4.1 of [5], the set

KC:=  ν ∈ Y : Z QT Z R2|ξ| 2 t,x(ξ)dxdt ≤ C  (4.6)

is compact in Y. Then, by the fundamental theorem for Young measures ([3], section 2), there exists ¯ νt,x∈ Y so that, up to a subsequence, lim δ→0 Z QT Z R2 J(t, x)f (ξ)dνt,xδ (ξ)dxdt = Z QT Z R2 J(t, x)f (ξ)d¯νt,x(ξ)dxdt (4.7)

for all continuous and bounded J : QT → R and f : R2 → R. We shall simply write νδ → ¯ν in

place of (4.7). By a simple adaptation of proposition 4.2 of [5], (4.7) can be extended to a function f : R2→ R such that f(ξ)/|ξ|2→ 0 as |ξ| → +∞.

In order to obtain (4.3), we need to prove that the limit Young measure ¯ν is a Dirac mass: ¯

νt,x = δu(t,x)¯ , for some ¯u ∈ L2(QT) and for almost every (t, x) ∈ QT. This is done using the

(8)

Definition 4.1. A Lax entropy-entropy flux pair for system (2.1) is a couple of differentiable func-tions (η, q) : R2→ R2 such that

(

ηr+ qp= 0

τ′(r)η

p+ qr= 0

. (4.8)

We show that Tartar’s equation holds for any two suitable entropy pairs (η, q) and (η′, q) to be

specified below and almost all (t, x) ∈ QT:

hηq′− η′q, ¯νt,xi = hη, ¯νt,xihq′, ¯νt,xi − hη′, ¯νt,xihq, ¯νt,xi, (4.9) where hf, ¯νt,xi := Z R2 f (ξ)d¯νt,x(ξ) (4.10)

for any measurable f . We employ the following argument due to Shearer [23].

Accordingly to Lemma 2 in [23], there exists a family of half-plain supported entropy-entropy fluxes (η, q) such that η and q are bounded together with their first and second derivatives. These are explicitly given as follows. We define z(r) :=R0rpτ′(ρ)dρ and we define the Riemann coordinates

w1= p + z, w2= p − z. We also pass from the dependent variables η, q to H,Q as follows:

η = 1 2(τ ′ )−1/4(H + Q) (4.11) q = 1 2(τ ′ )+1/4(H − Q) (4.12) so that (4.8) becomes ( Hw1= aQ Hw2= −aQ , (4.13) where a(w1− w2) = τ′′r w1− w2 2  8τ′r w1− w2 2 3/2. (4.14)

Then we fix ¯w1, ¯w2 ∈ R and we solve (4.13) with Goursat data given on the lines w1 = ¯w1 and

w2= ¯w2:

H( ¯w1, w2) = g(w2)

Q(w1, ¯w2) = 0,

(4.15)

where g is continuous and compactly supported. Then one can explicitly solve (4.13) and get

H(w1, w2) = g(w2) + ∞ X n=1 (Ang)(w1, w2) Q(w1, w2) = − Z w2 ¯ w2 a(w1− v)H(w1, v)dv, (4.16)

where the operator A acts on functions f ∈ L1

loc(R2) as follows: (Af)(w1, w2) = − Z w1 ¯ w1 Z w2 ¯ w2

a(v − w2)a(v − u)f(v, u)dudv. (4.17)

Finally, going back to η and q and using our assumptions on τ it is possible to show ([23], Lemma 2) that η and q are bounded, together with their first and second derivatives.

Now we have a suitable family of entropy-entropy flux pair, we use Tartar-Murat Lemma in order to derive Tartar’s equation (4.9). We evaluate (η, q) along the approximate solutions uδ and

compute the entropy production:

η(uδ)t+ q(uδ)x= δ  ηrrδx+ ηppδx  x− δ  ηrr(rδx)2+ ηpp(pxδ)2+ 2ηrprδxpδx  (4.18)

(9)

Since ηr and ηpare bounded and

√ δrδx,

δpδx are bounded in L2(QT), we have

lim δ→0δ  ηrrδx+ ηppδx  x= 0 in H −1 (QT), (4.19) while δ  ηrr(rxδ)2+ ηpp(pxδ)2+ 2ηrprδxpδx  L1(Q T) ≤ C (4.20) uniformly with respect to δ. Thus we have have obtained an equality of the form

η(uδ)t+ q(uδ)x= χδ+ ψδ,

where {χδ}

δ>0 lies in a compact set of H−1(QT) and {ψδ}δ>0 is bounded in L1(QT). Moreover,

since η and q are bounded, {η(uδ)t+ q(uδ)x}δ>0 is bounded in W−1,p(QT) for some p > 2.

Therefore, we can apply Tartar-Murat and the div-curl lemma (cf [9], Theorem 16.2.1 and Lemma 16.2.2) and obtain Tartar’s equation (4.9).

The final step is to use Tartar’s equation to prove that the support of the limit Young measure ¯

νt,xis a point. This is done in lemmas 4 to 7 of [23] and leads to the following

Proposition 4.2. There exists a ¯u ∈ L2(Q

T) such that ¯νt,x = δu(t,x)¯ for almost all (t, x) ∈ QT.

Moreover, uδ→ ¯u strongly in Lp(Q

T) for any p ∈ [1, 2).

4.1

Regularity

Proposition 4.3. For the function ¯u obtained in section 4,

¯

u ∈ L∞(0, T ; L2(0, 1)).

Proof. Since uδ→ ¯u in Lpstrong for p < 2, we can extract a subsequence {uδk

}k∈N that converges

pointwise to ¯u for almost all t and x. In particular, for almost all t, the sequence uδk(t, x) converges

for almost all x. Therefore, by Fatou lemma and Theorem 3.1, Z 1 0 |¯u(t, x)| 2 dx ≤ lim infk→∞ Z 1 0 |u δk (t, x)|2dx ≤ C (4.21) for almost all t ∈ [0, T ].

The proof is of next lemma is standard and therefore omitted.

Lemma 4.4. Let a(t) := τ−1τ (t)). Then, the solutions (rδ, pδ) of the viscous system (3.1) can be

written as follows: rδ(t, x) = a(t) + Z 1 0 Gδr(x, x′, t)(rδ0(x′) − a(0))dx′+ (4.22) + Z t 0 Z 1 0 Gδr(x, x′, t − t′)(∂x′pδ(t′, x′) − a(t′))dx′dt′ pδ(t, x) = Z 1 0 Gδp(x, x ′ , t)pδ0(x ′ )dx′+ Z t 0 Z 1 0 Gδp(x, x ′ , t − t′)∂x′τ (rδ(t′, x′))dx′dt′ (4.23) where the Gδ

r and Gδpare Green functions of the heat operator ∂t− δ∂xxwith homogeneous boundary

conditions: Gδr(1, x ′ , t) = ∂xGδr(0, x ′ , t) = 0 (4.24) Gδp(0, x′, t) = ∂xGδp(1, x′, t) = 0 (4.25) for all x′ ∈ [0, 1], t ≥ 0 and δ > 0.

(10)

The Green’s functions Gδr(x, x′, t) and Gδp(x, x′, t) are symmetric under the exchange of x and

x′. Moreover we have the following identities

∂xGδp(x, x ′ , t) = −∂x′Gδr(x, x ′ , t), (4.26) ∂xGδr(x, x′, t) = −∂x′Gδp(x, x′, t). (4.27)

Finally, Gδr and Gδpare subprobabilty densities, meaning

0 ≤ Z 1 0 Gδr(x, x′, t)dx′≤ 1 (4.28) 0 ≤ Z 1 0 Gδp(x, x′, t)dx′≤ 1 (4.29)

for any x, t and δ.

Remark. Although we will not use them, explicit expressions are available for the Gδ

r and Gδp, namely Gδp(x, x′, t) = X n odd e−tδλn sin√λnx  sin√λnx′  (4.30) Gδr(x, x ′ , t) = X n odd e−tδλn cos√λnx  cos√λnx′  , (4.31) with λn= n2π2 4 .

Proposition 4.5. For any φ ∈ C1([0, 1]), the application t 7→ Iφ(t) :=

Z 1 0

φ(x)¯u(t, x)dx (4.32)

is Lipschitz continuous. Consequently ¯u(t, ·) ∈ L2(0, 1) for all t ≥ 0.

Proof. We prove the statement for ¯p, as the proof for ¯r is similar. Furthermore, we prove the proposition only between 0 and t, as in the general case, say between t1 and t, it is enough to

replace the initial term pδ

0(x) with pδ(t1, x). We let Iφδ(t) := Z 1 0 φ(x)pδ(t, x)dx (4.33) and evaluate Iφ(t) − Iφ(0) = Z 1 0 Z 1 0 φ(x)Gδp(x, x′, t)pδ0(x′)dx′dx − Z1 0 φ(x)pδ0(x)dx+ (4.34) + Z t 0 Z 1 0 Z 1 0 φ(x)Gδp(x, x′, t − t′)∂x′τ (rδ(t′, x′))dxdx′dt′ = Z 1 0 φ(x) Z1 0 h Gδp(x, x ′ , t)pδ0(x ′ ) − pδ0(x) i dx′dx+ (4.35) + Z t 0 Z 1 0 Z 1 0 φ(x)∂xGδr(x, x′, t − t′)τ (rδ(t′, x′))dxdx′dt′+ Z t 0 Z1 0 φ(x)Gδp(x, 1, t − t′)¯τ (t′)dxdt′,

where we have used the symmetry of Gδ

pas well as the property ∂x′Gδ

p = −∂xGδr. The boundary term is estimated as Zt 0 Z 1 0 φ(x)Gδp(x, 1, t − t ′ )¯τ (t′)dxdt′ ≤ Cτ¯ Zt 0 Z 1 0 φ(x)Gδp(x, 1, t − t ′ )dx dt ≤ tCτ¯kφkL2, (4.36)

(11)

by Jensen’s inequality and since Gδpis a subprobability density.

We estimate the term involving τ as Z t 0 Z 1 0 Z1 0 φ(x)∂xGδr(x, x′, t − t′)τ (rδ(t′, x′))dxdx′dt′ = Z t 0 Z 1 0 Z 1 0 φ′(x)Gδ r(x, x′, t − t′)τ (rδ(t′, x′))dxdx′dt′ + (4.37) + φ(0) Z t 0 Z 1 0 Gδr(0, x′, t − t′)τ (rδ(t′, x′))dx′dt′ ≤ tkφ′kL2 1 t Z t 0 Z 1 0 Gδp(·, x′, t − t′)τ (rδ(x′, t′))dx′dt′ L2 + (4.38) + tkφkL∞ 1 t Z t 0 Z 1 0 Gδr(0, x′, t − t′)τ (rδ(t′, x′))dx′dt′ ≤ t kφ′kL2+ kφkL∞ 1 t Z t 0 Z 1 0 τ (rδ(x′, t′))2dx′dt′ 1/2 ≤ Ct (4.39) where C is independent of t and δ.

In order to estimate the first term of (4.35) we write Z 1 0 h Gδp(x, x′, t)pδ0(x′) − pδ0(x) i dx′=etδ∂xx − 1pδ0(x) (4.40) = Z tδ 0 ∂xxes∂xxpδ0(x)ds. (4.41) Therefore, we obtain Z 1 0 Z 1 0 h Gδp(x, x ′ , t)pδ0(x ′ ) − pδ0(x) i dx′dx ≤ Ztδ 0 Z 1 0 φ(x)∂xxes∂xxpδ0(x)dx ds (4.42) = Ztδ 0 Z 1 0 φ′(x)∂xes∂ xx pδ0(x)dx ds (4.43) ≤ kφ′ kL2 Z tδ 0 ∂xes∂xxpδ0 L2ds. (4.44)

Finally, since, by standard results on the heat equation, the application

s 7→ ∂xes∂xxpδ0 L2 (4.45) is decreasing, we have Z tδ 0 ∂xes∂xxpδ0 L2ds ≤ t √ δk√δ∂xpδ0kL2 ≤ Ct, (4.46)

where we have used the assumption that {√δ∂xpδ0}δ>0is bounded in L2(0, 1). We have also assumed,

without loss of generality, δ ≤ 1.

Putting everything together, we have obtained Iφδ(t) − Iφδ(0) ≤ Ct (4.47)

for some constant C independent of t and δ. This leads to the conclusion after passing to the limit δ → 0.

(12)

5

Proof of theorem 2.1 and Clausius inequality

All is left to prove is that the function ¯u obtained in the previous section is to a weak solution of the hyperbolic system (2.1), in the sense of Section 2. Let ψ ∈ C1(QT) with ψ(t, 0) = 0 for all t ∈ [0, T ].

Then, for any t ∈ [0, T ] we have 0 = Z t 0 Z 1 0  ψpδs− ψτ(rδ)x− δψpδxx  dxds = Z 1 0 ψ(t, x)pδ(t, x)dx − Z 1 0 ψ(0, x)pδ0(x)dx+ (5.1) − Zt 0 Z 1 0  ψspδ− ψxτ (rδ) − δψxpδx  dxds − Z ∞ 0 ψ(t, 1)¯τ (t)dt.

where we have used the initial-boundary conditions τ (rδ(t, 1)) = ¯τ (t) and pδ

x(t, 1) = 0, pδ(0, x) =

0(x) as well as ψ(t, 0) = 0. Since pδ0 converges to p0 in L2(0, 1), we have

lim δ→0 Z 1 0 ψ(0, x)pδ0(x)dx = Z 1 0 ψ(0, x)p0(x)dx. (5.2)

Furthermore, Theorem 3.1 implies√δpδx∈ L2(QT), consequently

Rt 0 R1 0 δψxp δ xdxds vanishes as δ →

0. Moreover, (4.2) implies, along the subsequence that defines ¯u = (¯r, ¯p),

lim δ→0 Z t 0 Z 1 0 ψtpδdxds = Z t 0 Z 1 0 ψspdxdt,¯ (5.3)

while by (4.3) we have that

lim δ→0 Z t 0 Z 1 0 ψxτ (rδ)dxds = Zt 0 Z 1 0 ψxτ (¯r)dxds,

so that (1.6) is satisfied. The (1.5) is linear and it follows similarly. Proposition 5.1. The solution ¯u satisfies Clausius inequality

F(¯u(t)) − F(u0) ≤ W (t) (5.4)

for all t ∈ [0, T ], where W (t) = − Z t 0 ¯ τ′(s) Z 1 0 ¯ r(s, x)dx + ¯τ (t) Z 1 0 ¯ r(t, x)dx − ¯τ(0) Z 1 0 r0(x)dx. (5.5)

Proof. By Proposition 4.3, Corollary 3.1, and Lemma 4.5, we have, for all t ∈ [0, T ], Z1 0  ¯ p2(t, x) 2 + F (¯r(t, x))  dx ≤ lim inf k→∞ Z 1 0  (pδk)2(t, x) 2 + F (r δk(t, x))  dx (5.6) ≤ lim k→∞  F(uδk 0 ) − Z t 0 ¯ τ′(s) Z1 0 rδk (s, x)dxds + ¯τ (t) Z 1 0 rδk (t, x)dx − ¯τ(0) Z 1 0 rδk 0 (x)dx  (5.7) = F(u0) − Zt 0 ¯ τ′(s) Z 1 0 ¯ r(s, x)dxds + ¯τ (t) Z 1 0 ¯ r(t, x)dx − ¯τ(0) Z 1 0 r0(x)dx, (5.8)

where we have used the fact that uδ

0 converges to u0 in L2 strongly in order to conclude that

F(uδ0) → F(u0). Moreover, all the integrals are well defined, since the application

t 7→ Z1 0 ¯ r(t, x)dx is continuous.

Thanks to the Clausius inequality, the solutions we have constructed are natural candidates for being the thermodynamic entropy solution of the equation (3.1) and one can conjecture that such limit is unique.

(13)

Acknowledgments

This work has been partially supported by the grants ANR-15-CE40-0020-01 LSD of the French National Research Agency. We thank Olivier Glass for very helpful discussions. We also thank an anonomous referee, whose critics and comments helped us to improve the first version of this article.

References

[1] P Acquistapace and B Terreni. On quasilinear parabolic systems. Mathematische Annalen, 282(2):315–335, 1988.

[2] L Alasio and S Marchesani. Global existence for a class of viscous systems of conservation laws. Preprint, https://arxiv.org/abs/1902.02714v1, 2019.

[3] J M Ball. A version of the fundamental theorem for young measures. In PDEs and continuum models of phase transitions, pages 207–215. Springer, 1989.

[4] John M Ball and Gui-Qiang G Chen. Entropy and convexity for nonlinear partial differential equations, 2013.

[5] F Berthelin and J Vovelle. Stochastic isentropic Euler equations. to appear in Annales de l’ENS, 2017.

[6] S Bianchini and A Bressan. Vanishing viscosity solutions of nonlinear hyperbolic systems. Annals of Mathematics, 161:223–342, 2005.

[7] G-Q Chen and H Frid. Vanishing viscosity limit for initial-boundary value problems for con-servation laws. Contemporary Mathematics, 238:35–51, 1999.

[8] G-Q G Chen and M Perepelitsa. Vanishing viscosity limit of the navier-stokes equations to the euler equations for compressible fluid flow. Communications on Pure and Applied Mathematics, 63(11):1469–1504, 2010.

[9] C M Dafermos. Hyperbolic conservation laws in continuum physics, volume 325 of grundlehren der mathematischen wissenschaften [fundamental principles of mathematical sciences], 2010. [10] R Di Perna. Convergence of approximate solutions to conservation laws. Archiv Rational Mech.

Anal., 82:27–70, 1983.

[11] Lawrence Evans. A survey of entropy methods for partial differential equations. Bulletin of the American Mathematical Society, 41(4):409–438, 2004.

[12] J Fritz. Microscopic theory of isothermal elastodynamics. Archiv Rational Mech. Anal., 201(1):209–249, 2011.

[13] P Lin. Young measures and an application of compensated compactness to one-dimensional nonlinear elastodynamics. Transactions of the AMS, 329(1):377–413, January 1992.

[14] S Marchesani. Hydrodynamic limit for a diffusive system with boundary conditions. Preprint, https://arxiv.org/abs/1903.08576, 2019.

[15] S Marchesani and S Olla. In preparation.

[16] S Marchesani and S Olla. Hydrodynamic limit for an anharmonic chain under boundary tension. Nonlinearity, 31(11):4979–5035, 2018.

[17] F Murat. Compacit´e par compensation. Annali della Scuola Normale Superiore di Pisa-Classe di Scienze, 5(3):489–507, 1978.

[18] F Murat. Compacit´e par compensation: condition n´ecessaire et suffisante de continuit´e faible sous une hypothese de rang constant. Annali della Scuola Normale Superiore di Pisa-Classe di Scienze, 8(1):69–102, 1981.

[19] S Olla. Microscopic derivation of an isothermal thermodynamic transformation. In From Particle Systems to Partial Differential Equations, pages 225–238. Springer, 2014.

(14)

[20] F Otto. Initial-boundary value problem for a scalar conservation law. Comptes rendus de l’Acad´emie des sciences. S´erie 1, Math´ematique, 322:729–734, 1996.

[21] D Serre. La compacit´e par compensation pour les syst`emes hyperboliques non lin´eaires de deux ´equations `a une dimension d’espace. J. Math. Pures Appl., 65(4):423–468, 1986.

[22] D Serre and J Shearer. Convercenge with physical viscosity for nonlinear elasticity. Preprint, unpublished, 1994.

[23] J W Shearer. Global existence and compactness in Lp for the quasi-linear wave equation.

Communications in Partial Differential Equations, 19(11):1829–1878, 1994.

[24] L Tartar. Compensated compactness and applications to partial differential equations. In Nonlinear Analysis and Mechanics: Heriot-Watt symposium, volume 4, pages 136–212, 1979. [25] L Tartar. The compensated compactness method applied to systems of conservation laws. In

Systems of Nonlinear Partial Differential Equations, pages 263–285. Springer, 1983. Stefano Olla

CEREMADE, UMR-CNRS, Universit´e de Paris Dauphine, PSL Research University

Place du Mar´echal De Lattre De Tassigny, 75016 Paris, France olla@ceremade.dauphine.fr

Stefano Marchesani GSSI,

Viale F. Crispi 7, 67100 L’Aquila, Italy stefano.marchesani@gssi.it

Références

Documents relatifs

Actually, in Section IV, it will be shown that in the presence of a moderate white noise added to the Sivashinsky equation, the solution is much more often close to the most

In [25], the authors apply to moment- SOS hierarchy to solve optimal control problems of ordinary differential equations, and it was shown in [31] that it provides a sequence

Samy Skander Bahoura. A compactness result for solutions to an equation with boundary singularity... 2018.. A compactness result for solutions to an equation with

Mathematics in En- gineering, Science and Aerospace (MESA) , 2018, Vol.. This paper deals with the construction of nonlinear boundary conditions for multidimensional conservation

It is well-known that compactness arguments based on BV bounds for the existence of solutions for quasilinear hyperbolic conservation laws is limited to one-dimensional

The goal of this article is to study the exponential stability of 2 × 2 systems of conservation laws on a finite interval, by means of boundary feedbacks, in the context of weak

We consider the wave equation in a smooth domain subject to Dirichlet boundary conditions on one part of the boundary and dissipative boundary conditions of memory-delay type on

(ii) As in the setting of the Barles–Souganidis theorem [2], where the Dirichlet bound- ary condition is relaxed from its classical pointwise sense, and is understood in a